18.01.2015 Views

LiDFOB - Chemical & Biomolecular Engineering - North Carolina ...

LiDFOB - Chemical & Biomolecular Engineering - North Carolina ...

LiDFOB - Chemical & Biomolecular Engineering - North Carolina ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Article<br />

pubs.acs.org/JPCC<br />

Solvate Structures and Computational/Spectroscopic<br />

Characterization of Lithium Difluoro(oxalato)borate (<strong>LiDFOB</strong>)<br />

Electrolytes<br />

Sang-Don Han, † Joshua L. Allen, † Erlendur Jońsson, ‡ Patrik Johansson, ‡ Dennis W. McOwen, †<br />

Paul D. Boyle, § and Wesley A. Henderson* ,†<br />

† Ionic Liquids & Electrolytes for Energy Technologies (ILEET) Laboratory, Department of <strong>Chemical</strong> & <strong>Biomolecular</strong> <strong>Engineering</strong>,<br />

<strong>North</strong> <strong>Carolina</strong> State University, Raleigh, <strong>North</strong> <strong>Carolina</strong> 27695, United States<br />

‡ Department of Applied Physics, Chalmers University of Technology, SE-412 96, Gothenburg, Sweden<br />

§ X-ray Structural Facility, Department of Chemistry, <strong>North</strong> <strong>Carolina</strong> State University, Raleigh, <strong>North</strong> <strong>Carolina</strong> 27695, United States<br />

*S Supporting Information<br />

ABSTRACT: Lithium difluoro(oxalato)borate (<strong>LiDFOB</strong>) is a relatively new salt designed for<br />

battery electrolyte usage. Limited information is currently available, however, regarding the<br />

ionic interactions of this salt (i.e., solvate formation) when it is dissolved in aprotic solvents.<br />

Vibrational spectroscopy is a particularly useful tool for identifying these interactions, but only<br />

if the vibrational bands can be correctly linked to specific forms of anion coordination. Single<br />

crystal structures of <strong>LiDFOB</strong> solvates have therefore been used to both explore the<br />

DFOB − ...Li + cation coordination interactions and serve as unambiguous models for the<br />

assignment of the Raman vibrational bands. The solvate crystal structures determined include<br />

(monoglyme) 2 :<strong>LiDFOB</strong>, (1,2-diethoxyethane) 3/2 :<strong>LiDFOB</strong>, (acetonitrile) 3 :<strong>LiDFOB</strong>, (acetonitrile)<br />

1 :<strong>LiDFOB</strong>, (dimethyl carbonate) 3/2 :<strong>LiDFOB</strong>, (succinonitrile) 1 :<strong>LiDFOB</strong>, (adiponitrile)<br />

1 :<strong>LiDFOB</strong>, (PMDETA) 1 :<strong>LiDFOB</strong>, (CRYPT-222) 2/3 :<strong>LiDFOB</strong>, and (propylene carbonate)<br />

1 :<strong>LiDFOB</strong>. DFT calculations have been incorporated to provide additional insight into the<br />

origin (i.e., vibrational modes) of the Raman vibrational bands to aid in the interpretation of<br />

the experimental analysis.<br />

1. INTRODUCTION<br />

The difluoro(oxalato)borate (DFOB − ) anion, also known as<br />

oxalyldifluoroborate (ODFB − ), was first reported in a patent<br />

application filed in 1999 by Metallgesellschaft AG 1 and in a<br />

publication in 2006 by the U.S. Army Research Laboratory<br />

(ARL). 2 The lithium salt with this anion (<strong>LiDFOB</strong>) has<br />

garnered significant interest in recent years for Li-ion battery<br />

electrolytes. 3−39 Little is known at present, however, about the<br />

molecular-level ionic interactions present once this salt is<br />

dissolved in the aprotic solvents used for electrolytes. Such<br />

knowledge is necessary to understand the origin of electrolyte<br />

properties as solvent−salt mixtures consist of a diverse range of<br />

solvate species. 40,41<br />

The solvates present in electrolytes may be broadly classified<br />

as solvent-separated ion pair (SSIP), contact ion pair (CIP) or<br />

aggregate (AGG) solvates if the anions are coordinated to zero,<br />

one or more than one Li + cation, respectively. Vibrational<br />

spectroscopy (Raman, FTIR) is the predominant means for<br />

determine the extent of ionic association in electrolytes by<br />

examining the variation in anion vibrational band positions.<br />

When an anion forms one or more coordination bonds to Li +<br />

cations, the anion electron density distribution (and thus bond<br />

lengths and angles) changes, leading to vibrational band<br />

variation. For Raman spectroscopy, the variability of the solvate<br />

species present in solution, however, typically results in broad<br />

overlapping anion Raman bands which can be difficult to<br />

deconvolute correctly and thus properly interpret to distinguish<br />

the specific forms of anion...Li + cation coordination. Computational<br />

methods may be utilized to assign the band positions for<br />

varying forms of coordination. 42−46 Unfortunately, such studies,<br />

while extremely useful for understanding the anion vibrational<br />

bands, do not always provide a definitive validation for the<br />

assignments. It has previously been demonstrated, however,<br />

that crystalline solvates with known structures can be utilized as<br />

models for the spectroscopic characterization of different forms<br />

of ClO 4 − ...Li + cation coordination. 47−52 These analyses are of<br />

particular relevance as they indicate that many of the vibrational<br />

assignments previously published for the coordination found in<br />

LiClO 4 -based electrolytes are either incorrect or oversimplifications<br />

of the true ionic association state present in the<br />

electrolytes.<br />

Received: September 13, 2012<br />

Revised: February 1, 2013<br />

Published: February 5, 2013<br />

© 2013 American <strong>Chemical</strong> Society 5521 dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Although useful for discussions, the general classification of<br />

SSIP, CIP, and AGG solvates fails to fully account for the<br />

coordination complexity that may exist. This classification must<br />

therefore be further differentiated to include specific forms of<br />

anion coordination, examples of which are shown in Chart 1 for<br />

Article<br />

Chart 2. Structures and Acronyms of the Solvents Studied<br />

Chart 1. DFOB ‐ ...Li + Cation Coordination: (a) SSIP, (b)<br />

CIP-I, (c) CIP-II, (d) AGG-Ia, (e) AGG-Ib, (f) AGG-Ic, (g)<br />

AGG-Id, (h) AGG-IIa, (i) AGG-IIb, (j) AGG-IIIa, and (k)<br />

AGG-IIIb a<br />

a Other forms of coordination may also occur. Li + cations are colored<br />

black.<br />

the DFOB − anion. Each of these different forms of<br />

coordination results in a different vibrational spectral fingerprint.<br />

To explore this in depth, the crystal structures for ten<br />

new crystalline solvate structures with <strong>LiDFOB</strong> are reported<br />

here. The Raman spectroscopic characterization of these<br />

solvates is also provided, as well as the data for the salt<br />

<strong>LiDFOB</strong> and other <strong>LiDFOB</strong> solvates for which a structure is<br />

known or can be reasonably presumed. In addition, density<br />

functional theory (DFT) calculations for the uncoordinated<br />

(i.e., SSIP) DFOB − anion and the CIP-II (Chart 1) form of<br />

DFOB − ...Li + cation coordination are presented to elucidate the<br />

vibrational modes which contribute to the vibrational bands<br />

and their correlation with the spectra for the crystalline solvates.<br />

2. EXPERIMENTAL SECTION<br />

2.1. Materials. <strong>LiDFOB</strong> was synthesized by the direct<br />

reaction of excess boron trifluoride diethyl etherate (BF 3 -ether)<br />

with lithium oxalate (oxalic acid dilithium salt), both used asreceived<br />

from Sigma-Aldrich. The resulting salt was extracted/<br />

recrystallized several times from dimethyl carbonate (DMC) as<br />

a (DMC) 3/2 :<strong>LiDFOB</strong> crystalline solvate. This solvate was<br />

subsequently vacuum-dried at 105 °C for 48 h, yielding the<br />

high purity anhydrous <strong>LiDFOB</strong> salt free of solvent. 11<br />

The solvents used in the present study and their acronyms<br />

are noted in Chart 2. The ethylene glycol dimethyl ether or 1,2-<br />

dimethoxyethane (monoglyme or G1, anhydrous, 99.5%),<br />

diethylene glycol dimethyl ether or 2-methoxyethyl ether<br />

(diglyme or G2, anhydrous, 99.5%), ethylene glycol diethyl<br />

ether or 1,2-diethoxyethane (Et-G1, 98%), N,N,N′,N″,N″-<br />

pentamethyldiethylenetriamine (PMDETA, 99%),<br />

1,1,4,7,10,10-hexamethyltriethylenetetramine (HMTETA,<br />

97%), 1,4,7,10-tetraoxacyclododecane (12-crown-4 or 12C4,<br />

98%), 4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8]-<br />

hexacosane (CRYPT-222, 98%), acetonitrile (AN, extra dry,<br />

99.9%), succinonitrile (SN, 99%), adiponitrile (ADN, 99%),<br />

tetramethylene sulfone or sulfolane (TMS, 99%), propylene<br />

carbonate (PC, anhydrous, 99.7%) and dimethyl carbonate<br />

(DMC, anhydrous, ≥ 99%) were purchased from either Fisher<br />

Scientific or Sigma-Aldrich and used as-received. The water<br />

content of the solvents was verified to be negligible (


The Journal of Physical Chemistry C<br />

corrected using a multiscan averaging of symmetry equivalent<br />

data using SADABS. 54 The structures were solved by direct<br />

methods using the SIR92 54 or XS 55 programs. All nonhydrogen<br />

atoms were obtained from the initial solution. The<br />

hydrogen atoms were introduced at idealized positions and<br />

were allowed to ride on the parent atom. The structural models<br />

were fit to the data using full matrix least-squares based on F 2 .<br />

The calculated structure factors included corrections for<br />

anomalous dispersion from the usual tabulation. The structures<br />

were refined using the XL program from SHELXTL. 55<br />

Additional details are available in the Supporting Information<br />

(SI). Structures were drawn using Ortep-3 or Mercury 3.0<br />

software. Crystallographic information data files (CIFs) for the<br />

solvates are available in the SI and free of charge from the<br />

Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.<br />

uk/data_request/cif: CCDC 896911 (ADN) 1 :<strong>LiDFOB</strong>,<br />

896912 (AN) 1 :<strong>LiDFOB</strong>, 896913 (AN) 3 :<strong>LiDFOB</strong>, 896914<br />

(CRYPT-222) 2/3 :<strong>LiDFOB</strong>, 896915 (DMC) 3/2 :<strong>LiDFOB</strong>,<br />

896916 (Et-G1) 3/2 :<strong>LiDFOB</strong>, 896917 (G1) 2 :<strong>LiDFOB</strong>, 896918<br />

(PC) 1 :<strong>LiDFOB</strong>, 896919 (PMDETA) 1 :<strong>LiDFOB</strong> and 896920<br />

(SN) 1 :<strong>LiDFOB</strong>. Mercury freeware software to view the content<br />

of these files is available: www.ccdc.cam.ac.uk/free_services/<br />

mercury/downloads/.<br />

2.4. Elemental Analysis. The analysis of crystals for the<br />

(12C4) 2 :<strong>LiDFOB</strong> and (HMTETA) 1 :<strong>LiDFOB</strong> solvates was<br />

performed by Atlantic Microlab, Inc.C 18 H 32 O 12 F 2 BLi for<br />

(12C4) 2 :<strong>LiDFOB</strong>, calcd: C 43.57, F 7.66, H 6.50; found: C<br />

43.50, F 7.53, H 6.36 and C 14 H 34 O 4 F 2 BLiN 4 for (HMTE-<br />

TA) 1 :<strong>LiDFOB</strong>, calcd: C 44.46, N 14.81, F 10.05 H 9.06; found:<br />

C 44.45, N 14.78, F 9.86, H 8.39confirming these as the<br />

correct compositions. A similar analysis was not performed on<br />

crystals for the (G2) 2 :<strong>LiDFOB</strong> solvate as this solvate melts<br />

below ambient temperature (SI).<br />

2.5. Raman Spectroscopy. Raman vibrational spectra<br />

were collected with a Horiba-Jobin Yvon LabRAM HR VIS<br />

high resolution confocal Raman microscope using a 632 nm<br />

He−Ne laser as the exciting source and a Linkam stage for<br />

temperature control with a long distance 50× objective. An<br />

elemental Si reference (520.7 cm −1 ) was used for spectral axis<br />

calibration. The zero position of the laser line and the Si line<br />

position were aligned prior to the data collection for each<br />

sample. The samples were hermetically-sealed in the Linkam<br />

stage in the glovebox before transferring the stage to the<br />

spectrometer. Spectra were typically collected using a 20 s<br />

measurement time with 10 accumulations. The samples were<br />

cooled/heated at 5 °C min −1 and the spectra were collected<br />

from −100 to 60 °C at intervals of 20 °C. Raman spectra were<br />

processed using LabSpec software.<br />

2.6. Computational Details. The DFOB − anion and its<br />

various ion pairs with a Li + cation were constructed manually<br />

and geometry optimized using the 6-311+G* basis set<br />

employing two different DFT functionals: B3LYP 56−58 and<br />

M06−2X 59 the former to support backward compatibility<br />

(significant previous work has been done using only the B3LYP<br />

functional) and the latter to make use of the recent<br />

development of better functionals (M06−2X). Utilizing both<br />

serves as an internal calibration/verification of the results.<br />

Subsequently, the structures were all verified as energy minima<br />

by computing their Hessians and the Raman spectral data<br />

(frequencies, activities) were obtained from the partial third<br />

derivatives of the energy. In addition to the above calculations,<br />

all done for ionic species in vacuum, a solvent effect was added<br />

(with the structures reoptimized) using a solvent self-consistent<br />

Article<br />

reaction field (SCRF-SMD) methodology 60 with AN as the<br />

parametrized solvent. This was done to in some way account<br />

for the reduced ion−ion interactions in the crystalline and/or<br />

liquid environment, and thus to mimic the experimental<br />

analysis to some extent. As the DFOB − anion is able to adopt a<br />

multitude of different forms of Li + cation coordination, the<br />

same strategy as noted above was also employed to scrutinize<br />

various ion pairs for numerous DFOB − ...Li + cation combinations.<br />

All calculations were made using Gaussian 09. 61 To assist<br />

the comparison with the experimental data, artificial Raman<br />

spectra were created by convolution of the spectral data using a<br />

Lorentzian band shape and a fwhm of 5 cm −1 .<br />

3. RESULTS AND DISCUSSION<br />

Although vibrational spectroscopy is a powerful tool for<br />

delineating the anion interactions with Li + cations, limitations<br />

exist for transcribing such information into a robust understanding<br />

of the ionic association interactions in solution. This is<br />

evident from a scrutiny of the solvate crystal structures. Figure<br />

1 shows examples of two different forms of anion CIP<br />

Figure 1. DFOB − ...Li + cation coordination in the crystalline solvates:<br />

(a) (G1) 2 :<strong>LiDFOB</strong> and (b) (Et-G1) 3/2 :<strong>LiDFOB</strong> (Li-purple, O-red, B-<br />

tan, F-light green).<br />

coordination in the (G1) 2 :<strong>LiDFOB</strong> and (Et-G1) 3/2 :<strong>LiDFOB</strong><br />

crystalline solvates. Vibrational spectroscopy only indirectly<br />

provides information about the Li + cation interactions. Thus,<br />

for the (Et-G1) 3/2 :<strong>LiDFOB</strong> crystalline solvate, the spectroscopic<br />

data would indicate a mixture of CIP-I and CIP-II<br />

solvates. But, from the perspective of the Li + cation<br />

coordination, the structure instead consists of SSIP and AGG<br />

solvates. Similarly, the spectroscopic data for the (CRYPT-<br />

222) 2/3 :<strong>LiDFOB</strong> solvate correspond to CIP-I/CIP-II/AGG-Id<br />

coordination (three symmetrically independent anions), whereas<br />

the Li + cation coordination is SSIP/SSIP/AGG. Molecular<br />

dynamics (MD) simulations of electrolyte mixtures point to a<br />

further complexity. In the liquid mixtures, there is a distribution<br />

of fully solvated Li + cations and uncoordinated anions, ion pairs<br />

5523<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Article<br />

Figure 2. DFOB − ...Li + cation coordination in crystalline <strong>LiDFOB</strong> and the crystalline solvates studied: (a) AGG-IIIb <strong>LiDFOB</strong>, 11 (b) AGG-Ic<br />

(H 2 O) 1 :<strong>LiDFOB</strong>, 11 (c) CIP-I (TMS) 2 :<strong>LiDFOB</strong>, 16 (d) CIP-II (G1) 2 :<strong>LiDFOB</strong>, (e) CIP-I/CIP-II (Et-G1) 3/2 :<strong>LiDFOB</strong>, (f) AGG-Ia<br />

(DMC) 3/2 :<strong>LiDFOB</strong>, (g) AGG-Ib (SN) 1 :<strong>LiDFOB</strong>, (h) AGG-Ia (AN) 3 :<strong>LiDFOB</strong>, and (i) AGG-IIa (AN) 1 :<strong>LiDFOB</strong>.<br />

and aggregate clusters of ions which may be present. 40,41 These<br />

solvated aggregate clusters may contain one or more anions<br />

coordination to one Li + cation (i.e., CIP coordination), but the<br />

cluster itself may actually be quite large with many ions (i.e., the<br />

Li + cation may be bonded to additional anions). 40 Thus, the<br />

spectroscopic data may again not be indicative of the true state<br />

of ionic association. Caution should therefore be exercised in<br />

directly translating the information obtained from a spectral<br />

analysis of electrolytes into an interpretation of solvate<br />

distribution as doing so may be highly misleading. Considerable<br />

insight from such an analysis may be obtained, however, when<br />

the spectroscopic data are used in tandem with other<br />

methods. 40,41<br />

3.1. Solvate Structures and Li + Cation Coordination.<br />

Figures 2 and 3 show schematic illustrations of the DFOB − ...Li +<br />

cation coordination in the crystalline solvates. The structures<br />

for pure <strong>LiDFOB</strong>, (H 2 O) 1 :<strong>LiDFOB</strong> and (TMS) 2 :<strong>LiDFOB</strong> have<br />

been previously reported. 11,16 The remaining structures were<br />

determined as part of the present study (SI). The DFOB −<br />

anion is able to adopt a multitude of different forms of Li +<br />

cation coordination. All of the anions in the solvate structures<br />

have at least one of the carbonyl oxygen atoms coordinated to a<br />

Li + cation. The two carbonyl oxygen atoms may coordinate up<br />

5524<br />

to three Li + cations using the four electron lone-pairs available.<br />

Bidentate coordination of a Li + cation by both carbonyl oxygen<br />

atoms is found in seven of the structures: <strong>LiDFOB</strong>,<br />

(G1) 2 :<strong>LiDFOB</strong>, (Et-G1) 3/2 :<strong>LiDFOB</strong>, (ADN) 1 :<strong>LiDFOB</strong>,<br />

(SN) 1 :<strong>LiDFOB</strong>, (AN) 1 :<strong>LiDFOB</strong> and (CRYPT-<br />

222) 2/3 :<strong>LiDFOB</strong>. This suggests that this is a common feature<br />

for the anion...Li + cation interactions in the solid-state<br />

structures. Other anions such as bis(oxalatoborate)<br />

(BOB − ), 44 dicyanotriazolate (DCTA − or TADC − ) 45 and<br />

bis(trifluoromethanesulfonyl)imide (TFSI − ) 41 may also have<br />

a tendency to form bidentate coordination, but this appears to<br />

diverge from the behavior of anions such as ClO − 4 ,BF − 4 , and<br />

PF − 6 , which instead typically coordinate Li + cation in solid-state<br />

solvate structures and in solution (the latter from MD<br />

simulations) via monodentate coordination. 50,62 Only four of<br />

the structures contain Li + cations coordinated by fluorine<br />

atoms: <strong>LiDFOB</strong>, (ADN) 1 :<strong>LiDFOB</strong>, (CRYPT-222) 2/3 :<strong>LiDFOB</strong><br />

and (PC) 1 :<strong>LiDFOB</strong>. In all of these solvates, the fluorine atom<br />

coordination serves as a necessary bridge to weld the structure<br />

together. This suggests that while fluorine atom coordination to<br />

Li + cations may occur, it is much less prevalent than the<br />

carbonyl oxygen atom coordination.<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Article<br />

Figure 3. DFOB − ...Li + cation coordination in the crystalline solvates studied: (a) AGG-IIb (ADN) 1 :<strong>LiDFOB</strong>, (b) AGG-Ia (PMDETA) 1 :<strong>LiDFOB</strong>,<br />

(c) CIP-I/CIP-II/AGG-Id (CRYPT-222) 2/3 :<strong>LiDFOB</strong>, and (d) AGG-IIIa (PC) 1 :<strong>LiDFOB</strong>.<br />

3.2. Raman Band Assignments. The experimental Raman<br />

spectrum for the anhydrous <strong>LiDFOB</strong> crystalline salt is shown in<br />

Figure 4. To aid in the assignment of the peaks, the DFT<br />

computed vibrational modes for the uncoordinated DFOB −<br />

anion are shown in Table 1. Note that differences are to be<br />

expected between the spectra for the anions in the highly<br />

aggregated salt and the uncoordinated anions. The vibrational<br />

modes may be visualized (through animation) using the<br />

Gaussian 09 output files (.xyz format) and instructions<br />

provided in the SI.<br />

The two strong peaks found above 1700 cm −1 in Figure 4<br />

correspond to the CO stretching vibrationsantisymmetric<br />

and symmetric, respectively (Table 1 and Chart 3 which note<br />

that the DFT calculations predict too high a frequency for these<br />

by ∼100 cm −1 ). Judging from the structural data, these bands<br />

should be excellent probes for the DFOB − ...Li + cation<br />

interactions as these occur principally via the anion carbonyl<br />

groups. This was found, however, not to be the case. A<br />

summary of the symmetric CO stretching band positions for<br />

the crystalline solvates is given in the SI. In contrast to<br />

expectations, this analysis suggests that the band positions for<br />

these DFOB − anion carbonyl vibrational modes do not enable<br />

the facile discrimination of the different forms of anion<br />

coordination.<br />

Figure 4. Raman spectrum of <strong>LiDFOB</strong> (at 20 °C).<br />

5525<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Article<br />

Table 1. Calculated (Uncoordinated DFOB − from M06-2X) Vibration Mode Frequencies (cm −1 ), Raman Activities (Å 4 amu −1 )<br />

and IR Intensities (km mol −1 ), and Tentative Assignments (No Scaling) a<br />

vacuum<br />

SMD<br />

freq expt freq I Raman I IR freq I Raman I IR tentative assignments<br />

81 1.61 2.10 82 2.55 2.53 B centered torsion<br />

114 0.72 0 110 1.09 0 torsion of OC−CO + ring + F−B−F<br />

325 331 0.28 ∼0 328 0.38 ∼0 ring twisting + torsion of F−B−F<br />

360 334 2.22 0.09 338 3.46 0.63 symm bend OC−O in-plane + F−B−F bend out-of-phase<br />

390 363 1.60 11.49 365 2.55 25.11 symm bend OC−O in-plane + F−B−F bend in-phase<br />

373 0.05 7.34 371 0.06 15.52 ∂(C−C) + π(F−B)<br />

468 0.12 19.63 467 0.42 33.07 π symm C−C<br />

522 0.30 25.05 517 0.44 40.49 ∂(OC−O) × 2+ρ rock B−F<br />

615 605 2.24 14.92 608 3.15 19.20 ∂(O−C−C)<br />

635 621 2.33 0.68 619 3.85 1.00 B−F scissoring + O−B symm bend<br />

680 690 0.44 2.78 695 0.96 2.95 ring rotation<br />

723 728 9.55 7.41 739 15.53 9.95 ring breathing<br />

840 856 0.53 ∼0 854 1.25 0.01 bend C−C (out of plane) + displacements of O<br />

942 957 2.82 0.49 964 5.33 2.10 symm B−O stretch + C−C stretch + B−F stetch<br />

987 0.04 240.71 989 0.14 358.15 antisymm B−O stretch<br />

1119 0.30 785.88 1078 0.63 1288.27 symm B−F stretch<br />

1177 0.29 382.10 1113 0.47 534.47 antisymm B−F stretch<br />

1290 0.01 197.31 1278 0.18 264.56 in-plane C−C bending + antisymm C−O stretch<br />

1405 1397 6.39 478.92 1386 7.61 677.46 symm C−O stretch + C−C stretch<br />

1769 1870 13.75 172.05 1865 25.17 223.30 antisymm CO stretch<br />

1800 1906 19.19 713.11 1900 45.02 912.41 symm CO stretch<br />

a The term freq expt refers to the <strong>LiDFOB</strong> salt (Figure 4).<br />

Chart 3. Principal Bonds/Angles Contributing to the<br />

DFOB − Anion Vibrational Modes as Determined from DFT<br />

Calculations (See SI)<br />

The peak observed at 1405 cm −1 (Figure 4) is attributed to<br />

the combination band of the C−C and symmetric C−O<br />

stretching modes (Table 1 and Chart 3). The peak near 942<br />

cm −1 (Figure 4) corresponds to a combination of symmetric<br />

B−O, C−C, and B−F stretching modes (Table 1 and Chart 3).<br />

The very strong peak near 723 cm −1 (Figure 4), predicted by<br />

the computations to be the third strongest peak in the Raman<br />

spectra, is due to a ring-breathing mode (Table 1 and Chart 3).<br />

The more complex modes corresponding to the two or more<br />

peaks observed at ∼600−620 cm −1 (Figure 4) are most likely a<br />

combination of different O−C−C bending modes, but<br />

combined also with B−F scissoring from the other half of the<br />

DFOB − anion (Table 1 and Chart 3). Thus, this region is more<br />

difficult to use as a probe for specific DFOB − ...Li + cation<br />

interactions. Yet, this should, at the same time, be the region<br />

with bands which undergo a sizable shift in position due to any<br />

Li + ...anion fluorine atom coordination. It is important to also<br />

note that boron isotopes have a natural abundance of 19.9% 10 B<br />

and 80.1% 11 B. Thus, vibrational modes which involve the<br />

displacement of the boron nucleus will give two different bands.<br />

This adds a further level of complexity to the band analysis for<br />

the DFOB − anion.<br />

3.3. Raman Spectroscopic Analysis of Crystalline<br />

Solvates. The majority of the solvates with known crystal<br />

structures were characterized by Raman spectroscopy to<br />

determine the anion band position variation with varying<br />

coordination and temperature. The AGG-Ic (H 2 O) 1 :<strong>LiDFOB</strong><br />

solvate, however, was not analyzed as this solvate contains<br />

significant hydrogen bonding between the water and anions<br />

(which is expected to perturb the vibrational bands), nor was<br />

the CIP-I/CIP-II/AGG-Id (CRYPT-222) 2/3 :<strong>LiDFOB</strong> solvate<br />

analyzed as the crystals melt below ambient temperature and it<br />

was not possible to prepare the correct liquid composition (i.e.,<br />

dissolve the appropriate amount of <strong>LiDFOB</strong> salt directly in the<br />

solvent). Unfortunately, the anion bands near 945 and 1405<br />

cm −1 overlap significantly with the peaks from many common<br />

solvents. The bands in the 600−650, 705−725, and 1780−1820<br />

cm −1 region of the spectra, however, do not and these bands<br />

5526<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Article<br />

Figure 5. Raman spectra of the DFOB − anion vibrational band variation for the (speculative) SSIP crystalline solvates: (12C4) 2 :<strong>LiDFOB</strong>,<br />

(G2) 2 :<strong>LiDFOB</strong>, and (HMTETA) 1 :<strong>LiDFOB</strong> (bold spectra indicate that the solvate has melted at this temperature).<br />

were therefore analyzed in depth. Representative data are<br />

shown in Figures 5 and 6. The remaining sets of data are<br />

provided in the SI. A summary of the 705−725 cm −1 results for<br />

the DFOB − anion Raman band positions as a function of<br />

temperature for the crystalline <strong>LiDFOB</strong> solvates is shown in<br />

Figure 7.<br />

The data in Figure 5 are for speculative SSIP solvates as no<br />

crystal structures are yet known for a SSIP solvate with<br />

<strong>LiDFOB</strong>. Attempts were made to determine the crystal<br />

structures of the (12C4) 2 :<strong>LiDFOB</strong>, (G2) 2 :<strong>LiDFOB</strong>, and<br />

(HMTETA) 1 :<strong>LiDFOB</strong> solvates, but these efforts were<br />

unsuccessful due to the very energetic solid−solid phase<br />

transitions which occur for these solvates at low temperature<br />

(SI) and the presence of significant disorder within the solvate<br />

structures in the higher temperature phases. There are no<br />

reported crystal structures yet known for solvates with<br />

HMTETA and LiX salts, but (12C4) 2 :LiX 63−67 and<br />

(G2) 2 :LiX 50,68−71 solvates are well-knownall of which have<br />

uncoordinated anions (i.e., SSIP solvates). The band positioned<br />

near 710 cm −1 for all three speculative SSIP solvates remains<br />

5527<br />

relatively fixed over the entire temperature range (Figures 5 and<br />

7). In contrast, multiple bands are present near 620 cm −1 which<br />

differ for the three solvates, as do the bands near 1760 and 1800<br />

cm −1 (Figure 5).<br />

The data in Figure 6 correspond to the CIP solvates with<br />

known crystal structures. There is rather poor conformity in the<br />

band positions for all three regions of the spectra for these<br />

solvates. The symmetric CO band near 1800 cm −1 (1906<br />

cm −1 from the calculations) also includes the C−O bonds<br />

(Chart 3). A comparison of the bond lengths/angles for the<br />

CIP-I anions (SI) suggests that these have a very similar<br />

geometry. If the bands near 1760 and 1796 cm −1 in the<br />

spectrum for the (Et-G1) 3/2 :<strong>LiDFOB</strong> solvate correspond to the<br />

CIP-I anion (rather than the CIP-II anion), then there is very<br />

good agreement with the same bands in the (TMS) 2 :<strong>LiDFOB</strong><br />

spectrum (Figure 6), but there is then poor agreement between<br />

the solvates for the bands near 620 cm −1 and 710 cm −1 (Figure<br />

6).<br />

The summary of the data for the ring breathing vibrational<br />

mode near 710 cm −1 shown in Figure 7 suggests that it will be<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Article<br />

Figure 6. Raman spectra of the DFOB − anion vibrational band variation for the CIP crystalline solvates: (TMS) 2 :<strong>LiDFOB</strong>, (G1) 2 :<strong>LiDFOB</strong> and (Et-<br />

G1) 3/2 :<strong>LiDFOB</strong> (bold spectra indicate that the solvate has melted at this temperature).<br />

quite difficult to accurately deconvolute the complex spectra for<br />

the DFOB − anion in the liquid phase to identify specific anion<br />

interactions. This is due to both the numerous forms of<br />

DFOB − ...Li + cation coordination which are possible and the<br />

variability in the band positions for any given form of anion<br />

coordination. This latter point is particularly true for the AGG<br />

solvates which have considerable variability in the anion<br />

geometry (SI). Given that each of the anion bands originates<br />

from the combination of multiple vibrations (Chart 3), it is not<br />

surprising that the anion geometry variation results in widely<br />

different spectra. In general, however, the band positions in the<br />

710−725 cm −1 range, especially when also considering the<br />

other regions of the spectra, do provide an indication of the<br />

relative fraction of SSIP, CIP, and AGG solvates (Figure 7). For<br />

example, upon melting, notable changes in the spectra occur for<br />

the (G1) 2 :<strong>LiDFOB</strong> and (Et-G1) 3/2 :<strong>LiDFOB</strong> solvates (Figure<br />

6). The spectra, which bear a close resemblance to the spectra<br />

for the melted (G2) 2 :<strong>LiDFOB</strong> solvate, suggest that the liquid<br />

melts obtained from the crystalline CIP solvates likely consist of<br />

a significant fraction of SSIP solvates (in addition to CIP<br />

solvates) with a lesser amount of AGG-I and perhaps AGG-II<br />

solvates. This is quite interesting given that melting of solvates<br />

often results in increased association (desolvation) of the<br />

solvate species rather than increased solvation (i.e., formation<br />

of SSIP solvates). 40,41<br />

Other forms of DFOB − ...Li + cation coordination than those<br />

found in the crystalline solvates (i.e., Chart 1) are possible/<br />

plausible in liquid electrolytes. DFT calculations identified five<br />

distinct CIP forms of coordination as energy minima (A−E,<br />

Chart 4), but only the <strong>LiDFOB</strong>-A ion pair was found to be<br />

strongly energetically favored for all four calculation methods<br />

employed (Table 2). Furthermore, as this type of bidentate<br />

DFOB − coordination via the carbonyl oxygen atoms is found in<br />

many of the crystalline solvates, as noted above, both the<br />

computational and experimental results suggest that there is a<br />

strong preference for this type of coordination. The vibrational<br />

modes for the <strong>LiDFOB</strong>-A ion pair are given in the SI, along<br />

with the Gaussian 09 output file for this ion pair.<br />

5528<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

Figure 7. Summary of the Raman band peak position for the DFOB −<br />

anion ring breathing vibrational band for various crystalline solvates<br />

(see SI for the assignments of the data to specific solvates).<br />

Chart 4. Stable Ion Pairs Obtained for <strong>LiDFOB</strong> from DFT<br />

Calculations a<br />

a While C and E look the same superficially, C has the Li + cation in the<br />

FBO plane and E has the cation in the oxalate plane.<br />

Table 2. Energies of the Calculated Ion-Pairs (Relative to<br />

<strong>LiDFOB</strong>-A)<br />

MO6−2X ΔE (kJ mol −1 ) B3LYP ΔE (kJ mol −1 )<br />

<strong>LiDFOB</strong>-A 0.0 <strong>LiDFOB</strong>-A 0.0<br />

<strong>LiDFOB</strong>-B 60.4 <strong>LiDFOB</strong>-B 59.5<br />

<strong>LiDFOB</strong>-C 51.0 <strong>LiDFOB</strong>-C 52.2<br />

<strong>LiDFOB</strong>-D 52.2 <strong>LiDFOB</strong>-D 51.6<br />

<strong>LiDFOB</strong>-E 62.6 <strong>LiDFOB</strong>-E 65.1<br />

MO6−2X-SMD ΔE (kJ mol −1 ) B3LYP-SMD ΔE (kJ mol −1 )<br />

<strong>LiDFOB</strong>-A 0.0 <strong>LiDFOB</strong>-A 0.0<br />

<strong>LiDFOB</strong>-B 33.5 <strong>LiDFOB</strong>-B 22.7<br />

<strong>LiDFOB</strong>-C 32.5 <strong>LiDFOB</strong>-C 35.9<br />

<strong>LiDFOB</strong>-D 28.4 <strong>LiDFOB</strong>-D 28.0<br />

<strong>LiDFOB</strong>-E 38.1 <strong>LiDFOB</strong>-E 37.4<br />

4. CONCLUSIONS<br />

Numerous crystal structures of <strong>LiDFOB</strong> solvates are reported<br />

which provide significant insight into the manner in which the<br />

DFOB − anion coordinates Li + cations. These solvates have<br />

been extensively characterized with Raman spectroscopy to link<br />

the vibrational bands with specific forms of DFOB − ...Li + cation<br />

coordination. This work has been complemented with DFT<br />

calculations to identify the origin (vibrational modes) of the<br />

DFOB − anion vibrational bands.<br />

5529<br />

■<br />

Article<br />

ASSOCIATED CONTENT<br />

*S Supporting Information<br />

Gaussian file for the DFOB − anion and <strong>LiDFOB</strong>-A ion pair to<br />

visualize anion vibrational modes; calculated Raman/IR<br />

vibrational information for the <strong>LiDFOB</strong>-A ion pair; differential<br />

scanning calorimetry (DSC) heating traces for the solvates;<br />

tables of bond lengths and angles for the CIP and AGG<br />

solvates; full Raman spectra for the AGG solvates; ion packing<br />

diagrams, additional details of the structure determination and<br />

crystallographic data (PDF) and crystallographic information<br />

files (CIFs) for the crystalline solvate structures (CIFs may be<br />

readily viewed with freeware programs such as Mercury or<br />

Jmol). This material is available free of charge via the Internet<br />

at ■http://pubs.acs.org.<br />

AUTHOR INFORMATION<br />

Corresponding Author<br />

*Tel: 919-513-2917. E-mail: whender@ncsu.edu.<br />

Notes<br />

The authors declare no competing financial interests.<br />

■ ACKNOWLEDGMENTS<br />

The authors wish to express their gratitude to the U.S.<br />

Department of Energy (DOE) Batteries for Advanced Transportation<br />

Technologies (BATT) Program which fully supported<br />

the experimental portion of this research under Award<br />

No. DE-AC02-05-CH11231. The Swedish Energy Agency is<br />

acknowledged for funding the computation portion of this<br />

research via a VR/STEM grant and SNIC for the allocation of<br />

computational resources. The authors also wish to thank the<br />

Department of Chemistry of <strong>North</strong> <strong>Carolina</strong> State University<br />

and the State of <strong>North</strong> <strong>Carolina</strong> for funding the purchase of the<br />

Apex2 diffractometer.<br />

■ REFERENCES<br />

(1) Tsujioka, S.; Takase, H.; Takahashi, M.; Sugimoto, H.; Koide, M.<br />

Electrolyte for Electrochemical Device. US Patent 6,506,516 B1, 2003<br />

(filed: June 7, 1999).<br />

(2) Zhang, S. S. An Unique Lithium Salt for the Improved Electrolyte<br />

of Li-Ion Battery. Electrochem. Commun. 2006, 8, 1423−1428.<br />

(3) Zhou, H.; Liu, F.; Li, J. Preparation, Thermal Stability and<br />

Electrochemical Properties of LiODFB. Appl. Mech. Mater. 2012,<br />

152−154, 1106−1111.<br />

(4) Yang, C.; Ren, Y.; Wu, B.; Wu, F. Formulation of a New Type of<br />

Electrolytes for LiNi 1/3 Co 1/3 Mn 1/3 O 2 Cathodes Working in an Ultra-<br />

Low Temperature Range. Adv. Mater. Res. 2012, 455−456, 258−264.<br />

(5) Aravindan, V.; Vickaman, P. Effect of Aging on the Ionic<br />

Conductivity of Polyvinylidenefluoride−Hexafluoropropylene<br />

(PVdF−HFP) Membrane Impregnated with Different Lithium Salts.<br />

Ind. J. Phys. 2012, 86, 341−344.<br />

(6) Hu, M.; Wei, J.; Xing, L.; Zhou, Z. Effect of Lithium<br />

Difluoro(oxalate) borate (<strong>LiDFOB</strong>) Additive on the Performance of<br />

High-Voltage Lithium-Ion Batteries. J. Appl. Electrochem. 2012, 42,<br />

291−296.<br />

(7) Dalavi, S.; Guduru, P.; Lucht, B. L. Performance Enhancing<br />

Electrolyte Additives for Lithium Ion Batteries with Silicon Anodes. J.<br />

Electrochem. Soc. 2012, 159, A642−A646.<br />

(8) Wu, X.; Wang, Z.; Li, X.; Guo, H.; Zhang, Y.; Xiao, W. Effect of<br />

Lithium Difluoro(oxalato)borate and Heptamethyldisilazane with<br />

Different Concentrations on Cycling Performance of LiMn 2 O 4 . J.<br />

Power Sources 2012, 204, 133−138.<br />

(9) Aravindan, V.; Gnanaraj, J.; Madhavi, S.; Liu, H.-K. Lithium-Ion<br />

Conducting Electrolyte Salts for Lithium Batteries. Chem.Eur. J.<br />

2011, 17, 14326−14346.<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

(10) Li, S.; Xu, X.; Shi, X.; Cui, X. Electrochemical Performance of<br />

Lithium Difluoro(oxalate)borate Synthesized by a Novel Method. Adv.<br />

Mater. Res. 2011, 197−198, 1121−1124.<br />

(11) Allen, J. L.; Han, S.-D.; Boyle, P. D.; Henderson, W. A. Crystal<br />

Structure and Physical Properties of Lithium Difluoro(oxalato)borate<br />

(<strong>LiDFOB</strong> or LiBF 2 O x ). J. Power Sources 2011, 196, 9737−9742.<br />

(12) Zugmann, S.; Fleischmann, M.; Amereller, M.; Gschwind, R. M.;<br />

Winter, M.; Gores, H. J. Salt Diffusion Coefficients, Concentration<br />

Dependence of Cell Potentials, and Transference Numbers of Lithium<br />

Difluoromono(oxalato)borate-Based Solutions. J. Chem. Eng. Data<br />

2011, 56, 4786−4789.<br />

(13) Zugmann, S.; Fleischmann, M.; Amereller, M.; Gschwind, R. M.;<br />

Wiemhöfer, H. D.; Gores, H. J. Measurement of Transference<br />

Numbers for Lithium Ion Electrolytes via Four Different Methods, a<br />

Comparative Study. Electrochim. Acta 2011, 56, 3926−3933.<br />

(14) Zhou, L.; Li, W.; Xu, M.; Lucht, B. Investigation of the<br />

Disproportionation Reactions and Equilibrium of Lithium Difluoro-<br />

(oxalato) Borate (<strong>LiDFOB</strong>). Electrochem. Solid-State Lett. 2011, 14,<br />

A161−A164.<br />

(15) Zugmann, S.; Moosbauer, D.; Amereller, M.; Schreiner, C.;<br />

Wudy, F.; Schmitz, R.; Schmitz, R.; Isken, P.; Dippel, C.; Müller, R.;<br />

et al. Electrochemical Characterization of Electrolytes for Lithium-Ion<br />

Batteries Based on Lithium Difluoromono(oxalato)borate. J. Power<br />

Sources 2011, 196, 1417−1424.<br />

(16) Allen, J. L.; Boyle, P. D.; Henderson, W. A. Lithium<br />

Difluoro(oxalato)borate Tetramethylene Sulfone Disolvate. Acta<br />

Crystallogr. 2011, E67, m533.<br />

(17) Amine, K.; Chen, Z.; Zhang, Z.; Liu, J.; Lu, W.; Qin, Y.; Lu, J.;<br />

Curtis, L.; Sun, Y.-K. Mechanism of Capacity Fade of MCMB/<br />

Li 1.1 [Ni 1/3 Mn 1/3 Co 1/3 ] 0.9 O 2 Cell at Elevated Temperature and<br />

Additives to Improve Its Cycle Life. J. Mater. Chem. 2011, 21,<br />

17754−17759.<br />

(18) Xu, M.; Zhou, L.; Hao, L.; Xing, L.; Li, W.; Lucht, B. L.<br />

Investigation and Application of Lithium Difluoro(oxalate)borate<br />

(<strong>LiDFOB</strong>) as Additive to Improve the Thermal Stability of Electrolyte<br />

for Lithium-Ion Batteries. J. Power Sources 2011, 196, 6794−6801.<br />

(19) Yang, L.; Markmaitree, T.; Lucht, B. L. Inorganic Additives for<br />

Passivation of High Voltage Cathode Materials. J. Power Sources 2011,<br />

196, 2251−2254.<br />

(20) Zygadło-Monikowska, E.; Florjanćzyk, Z.; Kubisa, P.; Biedroń,<br />

T.; Tomaszewska, A.; Ostrowska, J.; Langwald, N. Mixture of LiBF 4<br />

and Lithium Difluoro(oxalato)borate for Application as a New<br />

Electrolyte for Lithium-Ion Batteries. J. Power Sources 2010, 195,<br />

6202−6206.<br />

(21) Zhang, Z.; Chen, X.; Li, F.; Lai, Y.; Li, J.; Liu, P.; Wang, X. LiPF 6<br />

and Lithium Oxalyldifluoroborate Blend Salts Electrolyte for LiFe-<br />

PO 4 /Artificial Graphite Lithium-Ion Cells. J. Power Sources 2010, 195,<br />

7397−7402.<br />

(22) Fu, M. H.; Huang, K. L.; Liu, S. Q.; Liu, J. S.; Li, Y. K. Lithium<br />

Difluoro(oxalato)borate/Ethylene Carbonate + Propylene Carbonate<br />

+ Ethyl (Methyl) Carbonate Electrolyte for LiMn 2 O 4 Cathode. J.<br />

Power Sources 2010, 195, 862−866.<br />

(23) Li, J.; Xie, K.; Lai, Y.; Zhang, Z.; Li, F.; Hao, X.; Chen, X.; Liu, Y.<br />

Lithium Oxalyldifluoroborate/Carbonate Electrolytes for LiFePO 4 /<br />

Artificial Graphite Lithium-Ion Cells. J. Power Sources 2010, 195,<br />

5344−5350.<br />

(24) Moosbauer, D.; Zugmann, S.; Amereller, M.; Gores, H. J. Effect<br />

of Ionic Liquids as Additives on Lithium Electrolytes: Conductivity,<br />

Electrochemical Stability, and Aluminum Corrosion. J. Chem. Eng.<br />

Data 2010, 55, 1794−1798.<br />

(25) Huang, J.; Fan, L.-Z.; Yu, B.; Xing, T.; Qiu, W. Density<br />

Functional Theory Studies on the B-Containing Lithium Salts. Ionics<br />

2010, 16, 509−513.<br />

(26) Aravindan, V.; Vickraman, P.; Krishnaraj, K. Li + Ion Conduction<br />

in TiO 2 Filled Polyvinylidenefluoride-co-hexafluoropropylene Based<br />

Novel Nanocomposite Polymer Electrolyte Membranes with <strong>LiDFOB</strong>.<br />

Curr. Appl Phys. 2009, 9, 1474−1479.<br />

(27) Amereller, M.; Multerer, M.; Schreiner, C.; Lodermeyer, J.;<br />

Schmid, A.; Barthel, J.; Gores, H. J. Investigation of the Hydrolysis of<br />

Article<br />

Lithium Bis[1,2-oxalato(2-)-O,O′] Borate (LiBOB) in Water and<br />

Acetonitrile by Conductivity and NMR Measurements in Comparison<br />

to Some Other Borates. J. Chem. Eng. Data 2009, 54, 468−471.<br />

(28) Xiao, A.; Yang, L.; Lucht, B. L.; Kang, S.-H.; Abraham, D. P.<br />

Examining the Solid Electrolyte Interphase on Binder-Free Graphite<br />

Electrodes. J. Electrochem. Soc. 2009, 156, A318−A327.<br />

(29) Chen, Z.; Qin, Y.; Liu, J.; Amine, K. Lithium Difluoro(oxalato)-<br />

borate as Additive to Improve the Thermal Stability of Lithiated<br />

Graphite. Electrochem. Solid-State Lett. 2009, 12, A69−A72.<br />

(30) Gao, H.-Q.; Zhang, Z.-A.; Lai, Y.-Q.; Li, J.; Liu, Y.-X. Structure<br />

Characterization and Electrochemical Properties of New Lithium Salt<br />

LiODFB for Electrolyte of Lithium Ion Batteries. J. Cent. South Univ.<br />

Technol. 2008, 15, 830−834.<br />

(31) Aravindan, V.; Vickraman, P.; Krishnaraj, K. Lithium Difluoro-<br />

(oxalate)borate-Based Novel Nanocomposite Polymer Electrolytes for<br />

Lithium Ion Batteries. Polym. Int. 2008, 57, 932−938.<br />

(32) Kang, S.-H.; Abraham, D. P.; Xiao, A.; Lucht, B. L. Investigating<br />

the Solid Electrolyte Interphase Using Binder-Free Graphite Electrodes.<br />

J. Power Sources 2008, 175, 526−532.<br />

(33) Abraham, D. P.; Furczon, M. M.; Kang, S.-H.; Dees, D. W.;<br />

Jansen, A. N. Effect of Electrolyte Composition on Initial Cycling and<br />

Impedance Characteristics of Lithium-Ion Cells. J. Power Sources 2008,<br />

180, 612−620.<br />

(34) Chen, Z.; Liu, J.; Amine, K. Lithium Difluoro(oxalato)borate as<br />

Salt for Lithium-Ion Batteries. Electrochem. Solid-State Lett. 2007, 10,<br />

A45−A47.<br />

(35) Zhang, S. S. Electrochemical Study of the Formation of a Solid<br />

Electrolyte Interface on Graphite in a LiBC 2 O 4 F 2 -Based Electrolyte. J.<br />

Power Sources 2007, 163, 713−718.<br />

(36) Aravindan, V.; Vickraman, P. A Novel Gel Electrolyte with<br />

Lithium Difluoro(oxalato)borate Salt and Sb 2 O 3 Nanoparticles for<br />

Lithium Ion Batteries. Solid State Sci. 2007, 9, 1069−1073.<br />

(37) Liu, J.; Chen, Z.; Busking, S.; Amine, K. Lithium Difluoro-<br />

(oxalato)borate as a Functional Additive for Lithium-Ion Batteries.<br />

Electrochem. Commun. 2007, 9, 475−479.<br />

(38) Liu, J.; Chen, Z.; Busking, S.; Belharouak, I.; Amine, K. Effect of<br />

Electrolyte Additives in Improving the Cycle and Calendar Life of<br />

Graphite/Li 1.1 [Ni 1/3 Co 1/3 Mn 1/3 ] 0.9 O 2 Li-Ion Cells. J. Power Sources<br />

2007, 174, 852−855.<br />

(39) Zhang, S. S. A Review on Electrolyte Additives for Lithium-Ion<br />

Batteries. J. Power Sources 2006, 162, 1379−1394.<br />

(40) Seo, D. M.; Borodin, O.; Han, S.-D.; Ly, Q.; Boyle, P. D.;<br />

Henderson, W. A. Electrolyte Solvation and Ionic Association I.<br />

Acetonitrile-Lithium Salt Mixtures: Intermediate and Highly Associated<br />

Salts. J. Electrochem. Soc. 2012, 159, A553−A565.<br />

(41) Seo, D. M.; Borodin, O.; Han, S.-D.; Boyle, P. D.; Henderson,<br />

W. A. Electrolyte Solvation and Ionic Association II. Acetonitrile-<br />

Lithium Salt Mixtures: Highly Dissociated Salts. J. Electrochem. Soc.<br />

2012, 159, A1489−A1500.<br />

(42) Huang, W.; Frech, R.; Wheeler, R. A. Molecular Structures and<br />

Normal Vibrations of Trifluoromethane Sulfonate (CF 3 SO − 3 ) and Its<br />

Lithium Ion Pairs and Aggregates. J. Phys. Chem. 1994, 98, 100−110.<br />

(43) Gejji, S. P.; Suresh, C. H.; Babu, K.; Gadre, S. R. Ab Initio<br />

Structure and Vibrational Frequencies of (CF 3 SO 2 ) 2 N − Li + Ion Pairs. J.<br />

Phys. Chem. A 1999, 103, 7474−7480.<br />

(44) Holomb, R.; Xu, W.; Markusson, H.; Johansson, P.; Jacobsson,<br />

P. Vibrational Spectroscopy and Ab Initio Studies of Lithium<br />

Bis(oxalato)borate (LiBOB) in Different Solvents. J. Phys. Chem. A<br />

2006, 110, 11467−11472.<br />

(45) Scheers, J.; Niedzicki, L.; Zukowska, G. Z.; Johansson, P.;<br />

Wieczorek, W.; Jacobsson, P. Ion−Ion and Ion−Solvent Interactions<br />

in Lithium Imidazolide Electrolytes Studied by Raman Spectroscopy<br />

and DFT Models. Phys. Chem. Chem. Phys. 2011, 13, 11136−11147.<br />

(46) Johansson, P.; Beŕanger, S.; Armand, M.; Nilsson, H.; Jacobsson,<br />

P. Spectroscopic and Theoretical Study of the 1,2,3-Triazole-4,5-<br />

dicarbonitrile Anion and Its Lithium Ion Pairs. Solid State Ionics 2003,<br />

156, 129−139.<br />

(47) Grondin, J.; Talaga, D.; Lassègues, J.-C.; Henderson, W. A.<br />

Raman Study of Crystalline Solvates Between Glymes<br />

5530<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531


The Journal of Physical Chemistry C<br />

5531<br />

Article<br />

CH 3 (OCH 2 CH 2 ) n OCH 3 (n = 1, 2 and 3) and LiClO 4 . Phys. Chem.<br />

Chem. Phys. 2004, 6, 938−944.<br />

(48) Grondin, J.; Lassègues, J.-C.; Chami, M.; Servant, L.; Talaga, D.;<br />

Henderson, W. A. Raman Study of Tetraglyme−LiClO 4 Solvate<br />

Structures. Phys. Chem. Chem. Phys. 2004, 6, 4260−4267.<br />

(49) Henderson, W. A.; Brooks, N. R. Crystals from Concentrated<br />

Glyme Mixtures. The Single-Crystal Structure of LiClO 4 . Inorg. Chem.<br />

2003, 42, 4522−4524.<br />

(50) Henderson, W. A.; Brooks, N. R.; Brennessel, W. W.; Young, V.<br />

G., Jr. LiClO 4 Electrolyte Solvate Structures. J. Phys. Chem. A 2004,<br />

108, 225−229.<br />

(51) Henderson, W. A.; Brooks, N. R.; Brennessel, W. W.; Young, V.<br />

G., Jr. Triglyme−Li + Cation Solvate Structures: Models for<br />

Amorphous Concentrated Liquid and Polymer Electrolytes (I).<br />

Chem. Mater. 2003, 15, 4679−4684.<br />

(52) Henderson, W. A.; Brooks, N. R.; Young, V. G., Jr. Tetraglyme−<br />

Li + Cation Solvate Structures: Models for Amorphous Concentrated<br />

Liquid and Polymer Electrolytes (II). Chem. Mater. 2003, 15, 4685−<br />

4690.<br />

(53) SAINT and SADABS; Bruker AXS Inc.: Madison, WI, , 2009.<br />

(54) Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A.;<br />

Burla, M. C.; Polidori, G.; Camalli, M. SIR92A Program for<br />

Automatic Solution of Crystal Structures by Direct Methods. J. Appl.<br />

Crystallogr. 1994, 27, 435−436.<br />

(55) Bruker-AXS, XL version 2009.9; Bruker-AXS Inc.: Madison, WI,<br />

2009.<br />

(56) Vosko, S. H.; Wilk, L.; Nusair, M. Accurate Spin-Dependent<br />

Electron Liquid Correlation Energies for Local Spin Density<br />

Calculations: A Critical Analysis. Can. J. Phys. 1980, 58, 1200−1211.<br />

(57) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti<br />

Conelation Energy Formula into a Functional of the Electron Density.<br />

Phys. Rev. B 1988, 37, 785−789.<br />

(58) Becke, A. D. Density-Functional Thermochemistry. III. The<br />

Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652.<br />

(59) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals<br />

for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent<br />

Interactions, Excited States, and Transition Elements: Two<br />

New Functionals and Systematic Testing of Four M06-Class<br />

Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120,<br />

215−241.<br />

(60) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal<br />

Solvation Model Based on Solute Electron Density and on a<br />

Continuum Model of the Solvent Defined by the Bulk Dielectric<br />

Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113,<br />

6378−6396.<br />

(61) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;<br />

Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,<br />

B.; Petersson, G. A. et al. Gaussian 09, Revision B.01; Gaussian, Inc.:<br />

Wallingford CT, 2010.<br />

(62) Andreev, Y. G.; Seneviratne, V.; Khan, M.; Henderson, W. A.;<br />

Frech, R. E.; Bruce, P. G. Crystal Structures of Poly(Ethylene<br />

Oxide) 3 :LiBF 4 and (Diglyme) n :LiBF 4 (n = 1,2). Chem. Mater. 2005,<br />

17, 767−772.<br />

(63) Dillon, R. E. A.; Stern, C. L.; Shriver, D. F. Structural<br />

Comparisons of Fast Ion Conductors Consisting of Li[(CF 3 SO 2 ) 2 N]<br />

Complexes with Cryptands or Crown Ether. Solid State Ionics 2000,<br />

133, 247−255.<br />

(64) Lewis, R. A.; Wu, G.; Hayton, T. W. Stabilizing High-Valent<br />

Metal Ions with a Ketimide Ligand Set: Synthesis of Mn(NC t Bu 2 ) 4 .<br />

Inorg. Chem. 2011, 50, 4660−4668.<br />

(65) Hope, H.; Olmstead, M. M.; Power, P. P.; Xu, X. X-Ray Crystal<br />

Structures of the Diphenylphosphide and Arsenide Anions: Use of a<br />

Crown Ether to Effect Complete Metal Cation and Organometalloid<br />

Anion Separation. J. Am. Chem. Soc. 1984, 106, 819−821.<br />

(66) Gallucci, J. C.; Sivik, M. R.; Paquette, L. A.; Zaegel, F.; Meunier,<br />

P.; Gautheron, B. Two Derivatives of Lithium Isodicyclopentadienide:<br />

[(1,2,3,3a,7a-η)-4,5,6,7-Tetrahydro-4,7-methanoindenido]<br />

(N,N,N′,N′-tetramethylethylenediamine)lithium and Bis(1,4,7,10-<br />

tetraoxacyclododecane)lithium(1 + ) Bis[(1,2,3,3a,7a-η)-4,5,6,7-tetrahydro-4,7-methanoindenido]lithate(1<br />

− ). Acta Crystallogr. 1996, C52,<br />

1673−1679.<br />

(67) Liddle, S. T.; Clegg, W. A Homologous Series of Crown-Ether-<br />

Complexed Alkali Metal Amides as Discrete Ion-Pair Species:<br />

Synthesis and Structures of [M(12-crown-4) 2 ][PyNPh·PyN(H)Ph]<br />

(M = Li, Na and K). Polyhedron 2003, 22, 3507−3513.<br />

(68) Esterhuysen, M. W.; Raubenheimer, H. G. Lithium [Bis-<br />

{benzoyl-W(CO) 5 }(μ 2 -hydrogen)]−A Charge-Assisted H + -Bridged<br />

Organometallic Complex Prepared by Selective Li + Ion Replacement.<br />

Eur. J. Inorg. Chem. 2003, 2003, 3861−3869.<br />

(69) Pauls, J.; Iravani, E.; Koḧl, P.; Neumüller, B. Metallat-Ionen<br />

[Al(OR) 4 ] − als Chelatliganden für Übergangsmetall-Kationen. Z.<br />

Anorg. Allg. Chem. 2004, 630, 876−884.<br />

(70) Seneviratne, V.; Frech, R.; Furneaux, J. E.; Khan, M.<br />

Characterization of Crystalline and Solution Phases of Diglyme−<br />

LiSbF 6 . J. Phys. Chem. B 2004, 108, 8124−8128.<br />

(71) Palitzsch, W.; Boḧme, U.; Beyer, C.; Roewer, G. First Crystal<br />

Structure of a Cyclohexasilyl Transition-Metal Derivative,<br />

[(DIME) 2 Li][Mo(CO) 5 Si 6 Me 11 ] (DIME = Diethylene Glycol Dimethyl<br />

Ether). Organometallics 1998, 17, 2965−2969.<br />

dx.doi.org/10.1021/jp309102c | J. Phys. Chem. C 2013, 117, 5521−5531

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!